Miljø- og Fødevareudvalget 2015-16
MOF Alm.del
Offentligt
1643436_0001.png
Limnol. Oceanogr.,
59(5), 2014, 1679–1690
2014, by the Association for the Sciences of Limnology and Oceanography, Inc.
doi:10.4319/lo.2014.59.5.1679
E
Changes in the vertical distribution of primary production in response to
land-based nitrogen loading
Maren Moltke Lyngsgaard,
1,2,*
Stiig Markager,
1
and Katherine Richardson
2
1
Department
2
Center
of Bioscience, Aarhus University, Roskilde, Denmark
for Macroecology, Evolution and Climate, Danish Natural History Museum, University of Copenhagen, Copenhagen, Denmark
Abstract
Anthropogenic nitrogen (N)-loading has decreased significantly in the Baltic Sea Transition Zone over the past
two decades. We show that the vertical distribution of primary production (PP) changed as a function of land-
based N-loading using 1385 water column photosynthesis estimates, in which photosynthetic parameters were
determined both in the surface water layer and in the pycnocline-bottom layer (PBL) at six stations near the
Danish coast between 1998 and 2012. Total annual PP and surface layer PP (SPP) correlate positively with land-
based N-loading from Denmark (p
,
0.003). The percentage of annual PP occurring in the PBL (denoted as deep
primary production, DPP) varied annually between 6% and 30% (mean
5
17%). The absolute magnitude of the
DPP, as well as its relative proportion of total water column PP, correlates negatively with N-loading (p
,
0.009
and
p
,
0.0003, respectively). Thus, SPP decreases in response to decreased N-loading, while DPP increases.
Land-based N-loadings also correlate positively with the light attenuation coefficient (R
2
5
0.39,
p
,
0.05), which
may in part explain the response in DPP to changes in N-loading. DPP occurs in active phytoplankton
communities acclimated and/or adapted to low light and producing oxygen in the PBL water.
Primary production (PP) is an important factor in
structuring marine ecosystems, and changes in PP in
response to increased nutrient loading have been identified
as being responsible for symptoms of eutrophication in
coastal marine systems (Nixon 1995; Cloern 2001; Smith
2003). After recognizing the relationship between anthro-
pogenic nutrient loading and eutrophication, many coun-
tries have initiated programs to reduce nutrient enrichment
(Boesch 2002) with the expectation that marine PP will
respond to a reduction in land-based nutrient loading. In
some areas, enough data have now been collected to allow
researchers to examine the extent to which this expectation
has been realized.
The Baltic Sea Transition Zone (BSTZ), which com-
prises the Kattegat and the Belt Seas and thus forms the
connection between the Baltic Sea and the Skagerrak, is
such an area. It is a shallow, stratified, and temperate
marine system with dynamic hydrography. Surface salinity
varies in the region from 10 to 14 in the southern part of the
region and from 20 to 25 in the northern Kattegat
(Gustafsson 2000). Bottom water salinity generally varies
between 32 and 34, and an essentially permanent pycno-
cline is present. The area can be characterized as a frontal
system in which the low-saline surface water from the Baltic
Sea mixes with the more saline waters coming from the
Skagerrak. The system demonstrates clear estuarine circu-
lation where water transport is mainly driven by the water
level difference between the Arkona Sea and the Northern
Kattegat and, ultimately, by the freshwater surplus to the
Baltic Sea of 559 km
2
yr
21
(Savchuk 2005). Mixing with
surface water, especially in the Little and Great Belts (see
Fig. 1), ventilates the bottom water of the BSTZ (Bendtsen
et al. 2009).
* Corresponding author: [email protected]
The seasonal distribution of PP here is typical for
temperate coastal waters: Elevated production occurs in
association with the spring phytoplankton bloom, but peak
PP occurs during the summer months (Petersen and Hjorth
2010), as is also seen in other temperate estuarine regions
such as the Chesapeake Bay (Kemp and Boynton 1984).
Winter PP is low and limited by light availability.
Anoxia and hypoxia events in the BSTZ became more
widespread over the 20th century (Conley et al. 2007). As
these were believed to be a consequence of increases in
anthropogenic nutrient loading, legislation was established
in the late 1980s to control and reduce land-based nutrient
loading (Conley et al. 2002). Following the establishment of
this legislation, Denmark has maintained an extensive
marine monitoring program. The data set resulting from
this monitoring program provides a unique resource for
identifying relationships between PP and changing nutrient
loadings, which may be of potential relevance in under-
standing eutrophication responses in other coastal areas.
The purpose of this study was to examine these
monitoring data for evidence of a response in annual PP
in this region to changing nitrogen conditions. Because it is
well known that the BSTZ, as well as other seasonally or
semi-permanently stratified areas (Richardson et al. 2003;
Lehrter et al. 2009; Strom et al. 2010), are characterized by
a significant amount of annual PP taking place in
association with subsurface phytoplankton peaks (Cullen
1982; Richardson and Christoffersen 1991; Karlson et al.
1996), we chose not only to examine total water column PP
but also the seasonal and interannual variability in the
vertical distribution of PP. We hypothesized that PP
occurring in the pycnocline-bottom layer (referred to as
deep primary production, or DPP) may respond differently
than PP in the surface waters (SPP) to changes in land-
based nitrogen (N)-loading.
1679
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0002.png
1680
Lyngsgaard et al.
for the discrete depths at which sampling had been
conducted.
Data for total nitrogen (TN) loadings from Denmark to
the BSTZ were also taken from the MADS database. The
values are based on a three-dimensional Mike She (Windolf
et al. 2011) groundwater resource model validated on
measured TN concentrations in Danish streams. The
hydrological area chosen for this study covered the inner
Danish waters and ranged from the Kattegat Sea in the
north to the Belt Seas in the south.
The daily surface photosynthetic active radiation
(SPAR) was determined by averaging measurements made
continuously at several different localities in Denmark
within approximately 30–150 km of the sample locations.
Daily insolation was estimated using average SPAR
(between the different locations) calculated for every half
hour, and these average values were used to calculate the
24 h depth-integrated PP for all six stations. This data set is
not publically available.
Only data from 1998 through 2012 were used. This
period was chosen as the protocol for measuring
photosynthesis parameters was altered in 1998 so that
parameters were measured at two depths rather than one.
These two depths were an integrated ‘surface’ sample from
0 to 10 m (unless visual examination of the density profile
indicated a major density difference or pycnocline within
this depth range) collected using a hose inserted to the
desired depth or with Niskin bottles at 1, 5, and 10 m in
depth (or to the depth of the pycnocline) and the depth of
the deep chlorophyll maximum, defined as the depth with
the highest chlorophyll fluorescence and where fluores-
cence was greater than twice the average for the surface
layer (sample collected using Niskin bottle). Chl
a
concentrations were also determined in these two samples.
The protocol for the monitoring program prescribed that
if no deep chlorophyll maximum was encountered, the
second sample was taken at the depth of 2% surface light
penetration.
In addition to the two depths described above, Chl
a
was
analyzed at standard depths (every 5 m) throughout the
water column starting at 1 m below the surface and ending
1 m above the sediment. Chl
a
determinations were made
by filtering samples onto Whatman GF/F or GF 75
Advantec filters. Filters were extracted in ethanol (96%)
for 6–20 h and samples analyzed spectrophotometrically
according to the method described by Strickland and
Parsons (1972) and modified by Danish Standards (1986).
Photosynthetic carbon assimilation was estimated based
on the carbon-14 method modified (Markager 1998) after
Steemann Nielsen (1952). Photosynthesis vs. irradiance
curves (P vs. E curves) were calculated from incubations
made under artificial light (Osram HQI-T or high-pressure
halogen lamps), where the samples were incubated at seven
different light intensities for 2 h with metal grids providing
approximately 35% light attenuation between each bottle.
Thereafter, the samples were filtered (GF/F or GF 75
Advantec filters) and the carbon incorporation stopped
with acid (200
mL
0.1 mol L
21
HCl). The amount of
incorporated carbon-14 in the phytoplankton was determined
by liquid scintillation counting resulting in disintegrations per
Fig. 1. The Baltic Sea Transition Zone (BSTZ) and location
of the six stations.
Methods
Danish National Aquatic Monitoring and Assessment
Program—Data
from the 1998–2012 time period were
obtained from the database of the Danish National
Aquatic Monitoring and Assessment Program (MADS;
Conley et al. 2002). This database is publically available
(http://www2.dmu.dk/1_viden/2_Miljoe-tilstand/3_vand/4_
mads_ny/default_en.asp) and contains physical, chemical,
and biological data collected in BSTZ waters since the
1980s. Sampling and measurements were done by different
laboratories but followed a common set of technical
guidelines and procedures (Kaas and Markager 1998), and
inter-comparisons of results were frequently carried out by
the former Danish National Environmental Research
Institute (NERI), now Aarhus University. For the param-
eters used in this study, the inter-comparisons were
supervised by one of the authors (S.M.).
Six study stations were selected based on (1) a minimum
depth of 10 m and (2) the amount of data available. Study
station depths vary from 14 m (Aalborg Bight) to 51 m in
The Sound (Fig. 1). The four stations of intermediate
depths are located in Aarhus Bight, the Little Belt, and the
Great Belt (two stations). For each station, the following
data were extracted: conductivity, temperature, and depth
(CTD) profile data with 0.2 m resolution for the vertical
distribution of temperature, salinity, chlorophyll fluores-
cence, and photosynthetically active radiation (PAR; 4p
sensor). In addition, measurements of chlorophyll
a
(Chl
a)
and photosynthetic parameters (see below) were extracted
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0003.png
Baltic primary production and N-loading
1681
Fig. 2. Example of how parameters were derived for the primary production estimates: Profiles from 19 July 2010 in Aarhus Bight
of (a) density (r) and density difference m
21
(Dr
Dz
21
, kg m
24
). The starting depth of the pycnocline-bottom layer (PBL) was defined as
the first depth at which
Dr Dz
21
.
1 kg m
24
and, in the example here, was found at 6 m. (b) Interpolation of the chlorophyll-specific
photosynthetic parameters,
a
B
and P
B
. Actual determinations of these parameters at 1 and 15 m are indicated by black dots. (c) Primary
max
production, irradiance, and chlorophyll
a.
Irradiance is the sum over 24 h in 20 cm depth intervals.
minute (DPM). The equation used to calculate the carbon
assimilation in each sample was as follows:
P~
Tot CO
2
|
ð
DPM
tot
{DPM
dark
Þ|1:05|60
(DPM
start
|Light
minutes)
ð1Þ
where P is the measured production (mg C m
23
t
21
), Tot CO
2
is the total dissolved inorganic carbon concentration deter-
mined by Gran-titration, and Light minutes is the time (in
minutes) the phytoplankton have been exposed to light. The
14
C isotope discrimination factor is 1.05, and the value 60 is
used to calculate from minutes to hours. Light attenuation at
each station was determined by estimating the diffuse light
attenuation coefficient (K
d
) from the CTD profile using a
deck sensor as a reference.
Data analyses—The division of the water column into
two layers:
Density (r) was calculated from salinity and
temperature profiles (Fofonoff 1985). These density pro-
files were used to calculate the depth separating the surface
from the layers below (i.e., the pycnocline-bottom layer
[PBL]). This was defined as being at the first vertical
density gradient of
.
1 kg m
24
(see Fig. 2a for an
example). For each day with a distinct pycnocline (during
which this density criteria was met), every depth in the
water column was assigned as being in one of two layers
(i.e., in the surface layer or in the PBL). This gave the
possibility of considering the PP occurring in the surface
layer (surface primary production, SPP) and below the
surface (i.e., in the PBL [deep primary production, DPP]),
respectively. When the density criteria of
Dr/Dz
.
1 kg m
24
was not met in the water column, the total water column
PP was considered as being SPP.
Two hundred fifty-six visual inspections of density
profiles were made to confirm that the density criterion
was effective at determining the starting depth of the PBL.
The depths found by visual inspection matched the depths
found by the criterion in 90% of the cases. The density
criterion chosen (i.e., 1 kg m
24
) overestimated the starting
depth of the PBL (i.e., identified a deeper [1–2 m] depth
than identified visually) in the remaining 10%. Thus, our
estimate of the fraction of DPP as a percentage of total
water column PP (see below) is conservative.
Estimating Chl
a
from fluorescence—Fluorescence
per
unit of Chl
a
changed systematically with depth in the data
set. Therefore, a fluorescence factor (F
Chl
5
F/[Chl]) was
calculated whereby the chlorophyll concentration recorded
in the discrete sample (Chl) and the fluorometer measure-
ment from the corresponding depth (F) were related.
Values for F
Chl
between sampling depths were assigned by
linear interpolation. This was done with 0.2 m resolution
and the resulting profile of F
Chl
(z) used to estimate a
continuous Chl
a
profile (Chl(z)
5
F(z)/F
Chl
(z)) (see Fig. 2c
for an example of an estimated Chl
a
profile).
Estimating PP through the water column—To
estimate
water column PP, chlorophyll-specific photosynthesis rates
were calculated from the P vs. E curves and Chl
a
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0004.png
1682
Lyngsgaard et al.
Table 1.
Number of depth-integrated primary production measurements per year at the six sampling locations. In all, there were
1385 primary production measurements from year 1998 to 2012.
1998 1999
Aalborg Bight
Aarhus Bight
Little Belt
Great Belt 1
Great Belt 2
The Sound
26
49
12
49
18
2000
25
22
48
21
2001
25
48
17
2002
25
21
20
39
22
2003
19
28
48
23
2004
15
24
23
49
22
2005
17
25
24
23
45
22
2006
18
29
22
42
19
2007
20
27
23
2008
24
25
19
18
2009
2010
16
20
19
2011
18
19
19
20
2012
18
20
17
19
concentrations for the two samples, according to the method
of Markager et al. (1999). A light matrix representing the
light intensity at 0.2 m intervals throughout the water column
and hourly intervals over the entire day was constructed
using the attenuation coefficient and the surface light. This
light matrix and the photosynthetic characteristics derived
from the P vs. E curves (i.e.,
a,
P
max
, and offset (c) for the two
sampling depths normalized to Chl
a
concentrations [a
B
,
P
B
, and c
B
]) were combined to estimate daily PP at 0.2 m
max
intervals throughout the water column for all stations except
the Great Belt 1 Sta., at which only P vs. E curve parameters
from the surface layer were available. The P vs. E parameters
were obtained by a non-linear fitting procedure (Statistical
Analysis System, SAS, version 9.4) on the carbon uptake
from Eq. 2 divided by the chlorophyll concentration. The P
vs. E model was taken from Webb et al. (1974), modified by
the inclusion of an offset (c), thus:
!!
B
a
E
P
B
~P
B
zc
ð2Þ
max
1{exp
{
B
P
max
where
a
B
is the slope for the light-limited part of the P vs. E
curve (g C g
21
Chl
a
m
2
mol
21
), and P
B
the light-saturated
max
photosynthetic rate (g C g
21
Chl
a
h
21
). The offset (c) was
included in order to avoid a bias in
a
B
as a result of the
fact that
14
C-uptake rates are always positive.
See
also
Markager et al. (1999) for details about the curve fitting
procedure.
It was assumed that the P vs. E parameters from the
surface sample represented an average value for the entire
surface layer, and these values were therefore extracted from
the surface to the starting depth of the PBL (see Fig. 2 for an
example). The reason for using the same P vs. E parameter
values throughout the surface layer was that turbulence
within this layer was assumed to be too high to allow further
depth-specific photo adaptation (Lewis et al. 1984). Turbu-
lence in the PBL was assumed to be lower, thus allowing more
variation in the P vs. E parameter values with depth.
Therefore, these were interpolated from the starting depth
of the PBL (5 surface values) to the second sampling depth.
From the second sampling depth and downward photosyn-
thetic parameters were assumed to be constant (Fig. 2b). The
volumetric PP at each 0.2 m interval, P
vol
, (mg C m
23
d
21
)
was calculated from Eq. 3 with the following parameters:
P
vol
5
P
B
[Chl],
a
vol
5
a
B
[Chl], and c
vol
5
c
B
[Chl].
max
max
!!
PAR
P
vol
~P
vol
1{exp
{a
vol vol
ð3Þ
zc
vol
max
P
max
where PAR is the product of surface irradiance and the PAR
fraction left at a given depth [PAR fraction
5
(1
2
reflection)
exp (2K
d
(depth
2
0.05))]. The surface reflection was
assumed to be 6%.
In this manner, estimates of total water column PP as
well as the contribution made by DPP to the total were
made for the 1385 d of sampling. Data availability varied
between stations and, for each of the six stations, ranged
from 3 to 14 yr, resulting in a final time series of 15 yr
(Table 1). The sampling frequency at each station ranged
from 12 to 49 measurements per year.
The average annual values of PP were calculated for all
six stations. Long-term average monthly values were
calculated for each station (including all available years),
and these were used to fill months during which there were
no measurements. If data for entire years were missing,
these years were excluded from the analysis. This procedure
allowed us to examine the temporal development in PP
without the influence of missing values in the data set. We
used the same approach with respect to missing values for
the P vs. E parameters and the light attenuation coefficient
(K
d
) when analyzing their interannual and seasonal
variations.
Statistical analyses—Multiple
linear regressions were
constructed to examine the degree to which the estimated
PP could be explained by irradiance and N-loading.
Annual PP, annual DPP (g C m
22
yr
21
), DPP in
percentage of total, annual land-based N-loading to the
BSTZ from Denmark (10
3
kg of nitrogen), irradiance, and
the light attenuation coefficient (K
d
) were normalized to
their mean values over the study period. In this way the
unit of the coefficients for the regressions performed
becomes the percentage change in dependent variable
(PP, SPP, DPP, DPP in
%,
and K
d
) per percentage change
in forcing variables (N-loading and surface irradiance).
The annual land-based N-loading was calculated on a
monthly basis and summed over a period of consecutive
months. When relating PP parameters to N-loading, we
first regressed the relationships between the parameters for
a given calendar year against total N-loading in the same
calendar year. In recognition of the fact that loading might
affect PP differently over the season and that there likely
will be a lag between loading and PP response, we then
tested different periods for loading against the annual PP
parameters. A systematic approach was used in choosing
the periods to be tested. We started with 24 months
covering the whole calendar year before and the year of the
given annual PP. We then excluded and included months so
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0005.png
Baltic primary production and N-loading
Table 2.
Statistics for multiple linear regressions between primary production and the diffuse light attenuation coefficient (K
d
) vs. land-based N-loading from Denmark
and surface irradiance. The time period extends from 1998 to 2012. In addition, the temporal development in K
d
is analyzed from 1990 to 2012. Both dependent and
independent variables are scaled to the mean value
3
100; thus, the unit of the coefficients becomes
%
change in dependent variable per
%
change in the independent variable
(i.e.,
%
change in PP/% change in N-loading). Coefficients and intercepts are given
6
standard errors, and values in brackets are
p-values
for significance. Finally, each
dependent parameter is tested against time (annual change). Significant changes (p
,
0.05) are indicated in bold type.
1683
0.9960.64(0.15)
4.4±1.96(0.0006)
1.48±0.61(0.035)
3.2±0.66(0.0004)
Not significant
Not significant
Coefficient (C
2
)
for surface
irradiance
that all possible periods having a minimum length of
2 months were tested. In all, 276 different periods of N-
loading were tested against annual PP, SPP, and DPP. The
periods giving the highest explanatory power (R
2
) are
shown in Table 2.
The irradiance values used in the statistical analyses
represent the actual SPAR at the surface from April to
December as in all cases this period was that with highest
predictive value. A similar procedure to the one described
above was applied when the N-loading was considered in
relation to K
d
.
An analysis of the temporal development in the depth
separating the surface and PBL was carried out on
normalized depths (i.e., to the average depth of the PBL
for the individual station). When the normalized starting
depth of PBL was calculated for each station, a linear
regression was used on the normalized values as a function
of years per station and then for an average of the six
stations.
0.69
0.72
0.56
0.80
0.27
0.39
R
2
260±70
2276±98
292667
2144±68
89±5
90±4
Intercept
22.0(0.09)
3.6(0.08)
21.0(0.33)
3.6(0.026)
20.2(0.66)
20.9(0.017)
Annual
change,
%
Period for sum
of surface
irradiance
Results
Apr–Dec
Apr–Dec
Apr–Dec
Apr–Dec
Surface primary production
Deep primary production
Total water column production
Deep primary production in
%
of total
Diffuse attenuation coefficient (K
d
; 1998–2012)
Diffuse attenuation coefficient (K
d
; 1990–2012)
* Indicates months in the year before the calendar year for which primary production is calculated.
Dependent variable
Total and vertical distribution of PP—The
mean annual
PP from 1998 to 2012 for the six stations was 189, with
an interannual variation in the annual mean of
6
17.2
g C m
22
yr
21
. The lowest production estimates were found
in Aalborg Bight, Aarhus Bight, and The Sound (155–
175 g C m
22
yr
21
), while the production in Great and Little
Belts was higher (209–226 g C m
22
yr
21
; Fig. 3). The
higher values at these stations likely reflect nutrient inputs
from pycnocline-bottom to surface waters through more
intense mixing in the narrow straits (Lund-Hansen et al.
2008). The highest water column PP was found during
summer, and the seasonal pattern in production largely
followed irradiance and temperature (data not shown).
When all stations in the entire study period are
considered, the DPP, on average, contributed 17% to the
annual PP. The contribution of the DPP to the total water
column production varied from 6% in the Little Belt to
30% in Aalborg Bight (Fig. 3). The rest of the stations
exhibited contributions of DPP to annual production of
between 13% and 23% (Fig. 3). The highest average
monthly contribution from DPP to total was observed in
Aalborg Bight in May, where the DPP contributed 48% to
total water column production (Fig. 3c). DPP was low or,
in some cases, nonexistent from October to February, when
the phytoplankton community was well mixed throughout
the water column and surface irradiance was too low to
support production in the PBL (Fig. 3). Overall, the
importance of DPP was slightly higher in summer from
April to August (8–40% of summer production) than for
the year as a whole.
Photosynthetic parameters—The
seasonal pattern of the
light intensity at which photosynthesis initially is saturated,
I
k,
(P
max
/a,
mmol
photons m
22
s
21
) is shown in Fig. 4a as
the average for all six stations. Samples from the surface
layer show increasing I
k
values from January (68
mmol
photons m
22
s
21
) to a peak in August (128
mmol
photons
m
22
s
21
), followed by a decline to 78
mmol
photons m
22
s
21
Period for sum
of N-loading
Coefficient (C
1
) for
N-loading from
Denmark
Feb–Sep
Mar*–Feb
Feb–Sep
Mar*–Feb
Nov*–Mar
Nov*–Mar
0.61±0.12(0.0002)
20.66±0.21(0.0090)
0.43±0.11(0.0025)
20.73±0.14(0.0003)
0.10±0.05(0.0495)
0.09±0.03(0.0121)
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0006.png
1684
Lyngsgaard et al.
Fig. 3. Monthly average values for total water column primary production (PP) and deep PP (DPP) for six different stations in the
BSTZ (1998–2012). The annual PP and the contribution (in
%)
from DPP to PP are given in the upper right corner of each graph.
in December. For the deep samples, the seasonal variation
was smaller. I
k
increased from 62
mmol
photons m
22
s
21
in
May to 89
mmol
photons m
22
s
21
in November and then
dropped to 55
mmol
photons m
22
s
21
in January. A
t-test
showed significantly higher average monthly I
k
values in
the surface layer than in the PBL (t-test,
t
5
3.56, degrees
of freedom [df]
5
22,
p
,
0.01,
n
5
24). In addition to the
characteristic seasonal pattern seen for the I
k
values from
the surface layer, a significant and positive relationship was
found between I
k
values from the surface layer and the
daily surface irradiance (r
5
0.28,
p
,
0.0001,
n
5
1080).
Thus, the I
k
data indicate that the phytoplankton
communities in and below the surface layer are acclimated
to the prevailing light levels they encounter.
Not surprisingly, this conclusion is also supported by the
general patterns observed for maximum chlorophyll-specific
photosynthesis rate (P
B
), as this parameter controls a large
max
part of the variation in I
k
. Average monthly values
(including data from 15 yr) for P
B
showed significantly
max
higher values for phytoplankton populations found in the
surface layer (P
B
5
3.6
6
1.4 g C g
21
Chl
21
h
21
) than in
max
the PBL (P
B
5
2.3
6
0.4 g C g
21
Chl h
21
;
t-test, t
5
3.06,
max
df
5
22,
p
,
0.01,
n
5
24). An exception to this general
pattern was observed for the Sound, where P
B
values from
max
the two depths were similar.
Monthly averages of P
B
values in the surface layer for
max
all six stations for the 1998–2012 period exhibited a
seasonal pattern in which P
B
increased from 2.4 g C g
21
max
Chl h
21
in March to 6.4 g C g
21
Chl h
21
in August and
then decreased to 2.2 g C g
21
Chl h
21
in December (Fig. 4).
P
B
max
values from surface and PBL waters were not
significantly different (t-test,
t
5
1.83, df
5
734,
p
5
0.07,
n
5
736) between the months of December and
March. This can also be seen from the average monthly
values (Fig. 4). During this period, the water column was
often well mixed and the solar irradiance so low that
photosynthesis was likely light limited most of the time.
The general pattern in the chlorophyll-specific initial
slope (a
B
) of the P vs. E curves exhibited no significant
difference for phytoplankton populations found in the
surface layer and those found in the PBL (Fig. 4c). In
addition, the variability, expressed as the coefficient of
variance (CV), of
a
B
was lower than that of P
B
(CV
a
B
5
max
0.16 compared to CV for P
B
5
0.42). This is to be
max
expected, as
a
is equal to absorption times quantum yield
(Markager and Vincent 2001), and light absorption is
tightly coupled to the chlorophyll content in the cells
(Markager 1993). A weak seasonal pattern in
a
B
values was
detected that resembled that found for I
B
and P
B
. The
a
B
k
max
values increased from 10.2 g C g
21
Chl mol photons
21
m
2
in March to the peak value in August (14.2 g C g
21
Chl
mol
21
m
2
), and decreased thereafter to 8.2 g C g
21
Chl
mol
21
m
2
in December. Higher values were found in
January and February than in March and April.
PP and land-based N-loading—The
concentration of
nitrogen in land runoff from Denmark (mg N L
21
)
decreased significantly between 1998 and 2012 (linear
regression:
p
,
0.0001,
R
2
5
0.84,
n
5
15; Fig. 5),
primarily as a result of the establishment of water quality
regulation (Windolf et al. 2012). Although there were large
fluctuations in the absolute loading due to interannual
variation in runoff, the absolute delivery of nitrogen from
Denmark (10
3
kg N) also showed a significant decrease
over the study period (linear regression:
p
,
0.016,
R
2
5
0.37,
n
5
15;
see
Fig. 5).
Total PP, SPP, and DPP, as well as the ratio of SPP
and DPP to total PP, were all significantly correlated to
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0007.png
Baltic primary production and N-loading
1685
Fig. 5. Land-based annual N-loading (right y-axis) and land-
based annual N-loading normalized to runoff (left y-axis) from
1998 to 2012. The numbers are based on average values of Danish
N-loading to the study area.
Fig. 4. Seasonal variation of chlorophyll
a–specific
P vs. E
parameters: (a) the light intensity at which photosynthesis is
initially saturated, I
k
; (b) maximum chlorophyll-specific photo-
synthesis rate, P
B
; and (c) the initial slope of the P vs. E curve,
max
a
B
. Parameters are shown for 1–10 m (depth 1) and for depths in
the pycnocline-bottom layer (depth 2). The parameters are
average values of six stations in the Baltic Sea Transition Zone
(1998–2012), and the error bars indicate the variation in average
monthly values among stations.
land-based N-loading (Table 2). However, while SPP was
positively correlated to N-loading (see Fig. 6), the opposite
was true for DPP.
The relationship between SPP and N-loading can be seen
in Fig. 6, where both SPP and N-loading are calculated for
the calendar year. The relationship is significant, with a
p-
value
,
0.01 and a coefficient of 0.47% per
%
change in N-
loading. Thus, when N-loading changed by 1%, the SPP
changed by 0.47% in the same direction. It is, however,
doubtful that N-loading calculated over the calendar year is
the best descriptor in terms of relating N-loading to PP, as
N-loading late in the calendar year obviously cannot affect
the period of peak production during the same year.
Instead, it can be argued that nitrogen has a residence time
in the system. This would argue for including N-loading
from the previous year when considering PP in a given year,
especially given the fact that phytoplankton growth is only
believed to be N-limited during the summer period in the
region (Conley 1999). Following this line of reasoning, the
N-loading during the summer might be most important in
controlling PP. We therefore tested a number of different
loading periods extending back to January of the year prior
to the test year in relation to the PP variables.
A change in N-loading was the best predictor of changes
in SPP when the 8 months from February to September of
the study year were included (coefficient of 0.61;
p
,
0.001;
Table 2). Surface irradiance was also positively related to
SPP, but the coefficient was not significant. Figure 7 shows
the effect on the
R
2
value (explained variance) by varying
the timing of an 8 month loading period. All 8 month
periods with a midpoint from 01 November in the year
prior to the study year to 01 July of the study year yielded
significant (p
,
0.05) coefficients for the correlation
between SPP and N-loading. The value of the coefficient
varied from 0.42 to 0.61. This suggests that the N-loading
from the previous winter and during the period in which
phytoplankton activity is greatest are the most important
drivers of SPP. Calculating the N-loading over periods with
different lengths gave similar results.
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1643436_0008.png
1686
Lyngsgaard et al.
Fig. 6. Changes in annual primary production in the surface
layer as a function of changes in N-loading. Change in each
parameter is expressed as a percentage of the average value for the
study period (1998–2012) as a whole. A significant linear
relationship between the two parameters (p
,
0.01 and
R
2
5
0.41) is noted on the graph with a dashed line.
Fig. 7. An example of the relationships between periods over
which the N-loading is calculated as well as the resulting
R
2
values
and coefficients. The month is the midpoint of the period (an
asterisk indicates the year before the calendar year during which
primary production is calculated), and the period from Table 2 is
indicated in bold. The horizontal dashed line indicates when
the
p-value
is below 0.05. The example shown here is for
surface primary production vs. N-loading calculated over an
8 month period.
The DPP was strongly and positively correlated with
surface irradiance (coefficient
5
4.4%/%,
p
,
0.001) but
negatively related to the N-loading (Table 2). In this case,
the most significant relationship was found when N-
loading was calculated over a 12 month period starting in
March of the year prior to the study year to February of
the study year (midpoint between August and September of
the year prior to the study year). The coefficient to N-
loading was negative for all periods but was only significant
for 12 month periods with midpoints between August and
October of the year prior to the study year (Table 2), which
means that changes in DPP were best predicted by a period
of N-loadings occurring prior to the actual DPP.
Total PP was also significantly and positively related to
N-loading (summed from February to September of the
study year; i.e., similar to the case for SPP). However, the
relationship was weaker than the relationship between
loading and SPP. This can easily be explained as the total
water column PP is the sum of SPP and DPP and these two
layers exhibited opposite responses to N-loading.
The most significant model relating PP parameters to N-
loading was that describing the fraction of DPP (as a
percentage of total water column production) as a function
of N-loading. This indicates that N-loading potentially
redistributes PP in the water column so it shifts from the
surface layer to the PBL as loadings are reduced. This result
cannot be attributed to changes in the depth at which the
surface and PBL diverge as there was no significant temporal
development in the starting depth of the PBL when the
stations were analyzed by linear regression separately (p
.
0.05) as well as when they were averaged into one value
(normalized to the average depth for the study period as a
whole) per year (p
.
0.05). A possible explanation for the
identified negative relationship between DPP and N-loading
is the positive relationship found between N-loading and
light attenuation in the water column. We therefore also
tested the relationship between N-loading and the diffuse
attenuation coefficient (K
d
). A significant positive relation-
ship was found with a coefficient of 0.10% in K
d
per percent
change in N-loading (Table 2). K
d
data were available from
1990. Using this longer time series gave the same coefficient
but with a lower
p-value
(Table 2). There was a positive
relationship between the interannual variation (relative to
the long-term average) in K
d
and surface layer chlorophyll
(R
2
5
0.30,
p
,
0.05,
n
5
15). Thus, the ‘extra’ PP stimulated
by a higher N-loading increases the chlorophyll concentra-
tion in the surface layer, contributing to the drop in water
transparency.
Temporal development in biological and physical param-
eters—The
temporal trends of the studied parameters in
response to changes in land-based N-loading are masked by
the large interannual variability in runoff. Nevertheless, we
also regressed all of the parameters studied against year
(Table 2, last column). While the trends expected in light of
the overall reduction in N-loadings were found, two more
parameters showed a significant temporal development.
The K
d
decreased over time (for the time period from 1990
to 2012) and the fraction of DPP to total PP showed an
increase from 1998 to 2012 (see Table 2). No temporal
trends were found for PP, SPP, or DPP (as absolute carbon
values). Thus, over the 1998–2012 period, there has been a
significant vertical redistribution of PP in the BSTZ,
whereby PP has moved from the surface waters to the
PBL and water clarity has increased.
Discussion
PP in relation to land-based N-loading—This
study
demonstrates a redistribution of PP in the water column
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
Baltic primary production and N-loading
with decreasing land-based N-loading from Denmark,
whereby the proportion of PP taking place beneath the
surface layer increases with decreasing loading. This
finding is potentially important for the management of
eutrophication induced by human activities. From our
general understanding of eutrophication (Cloern 2001),
such a redistribution of PP in response to anthropogenic
eutrophication might have been predicted. However, to our
knowledge, this is the first time it has been demonstrated
for a coastal ecosystem on a decadal timescale. That it has
not been documented earlier with field observations is
likely due to the fact that there are few locations with
substantial time series of data describing phytoplankton
photosynthetic activity over a period with concurrent
reduction in nutrient loading.
The coefficient of the regression relating SPP and N-
loading (Table 2) indicates that SPP changes 0.61% per
percent change in N-loading (using load period from
February to September of the study year). The intercept for
this regression (intercept
5
60%) indicates that more than
half of the production is based on regenerated N or N from
other sources rather than land-based loading (see below).
In the case of the DPP, a 1% increase in land-based N-
loading leads to a 0.66% decrease in DPP. An interesting
consequence of the fact that opposite responses to the
reduction of N-loading are seen in the two depth layers is
that the total water column production is not as sensitive to
changes in N-loading as are the two individual layers when
examined independently. Thus, total water column PP may
not be the ideal parameter to use in assessing system
responses to changes in nutrient loading.
That we see a greater shift in the distribution of PP in the
water column than in total PP is reminiscent of the pattern
observed for more shallow aquatic systems, in which an
increase in N-loading leads to a redistribution of PP from
benthic macrophytes to phytoplankton (i.e., higher loading
results in the production occurring closer to the surface;
Borum and Sand-Jensen 1996; Krause-Jensen et al. 2012).
As no temporal development in the starting depth of the
PBL was recorded, the most likely explanation for the
patterns observed in both cases must be that higher loading
leads to a reduction in water clarity. This is supported by
the relationship found in this study between the N-loading
and K
d
. That lower
R
2
values are found for K
d
(0.27–0.39)
compared to PP (0.56–0.80) in relation to nutrient loading
agrees well with earlier studies, in which it has been
demonstrated that the relationship between N-loading and
ecosystem effects becomes weaker when the response of the
parameter is not directly coupled to the loading being
examined (Timmermann et al. 2010; Carstensen et al.
2011).
The best fits between land-based N-loading and PP were
found when loading was summed over periods other than
the calendar year for which the PP was calculated. For SPP
and PP, the period best describing the relationship to N-
loading (judged on the basis of the value of
R
2
) was
February to September of the study year, but the winter
months prior to this period were also shown to be
important (Fig. 7). Thus, loading during the period when
PP is high has a large effect on the PP occurring. In
1687
contrast, both the DPP and K
d
values were better described
by N-loading occurring farther back in time (Table 2).
Assuming that the negative coefficient for N-loading vs.
DPP derives, ultimately, from the positive effect on light
attenuation, both observations would suggest the occur-
rence of a lag period from when the loadings reach the
marine system until the effects are observed.
Light attenuation is governed by several processes and
components in the water column (i.e., chlorophyll and
accumulated organic matter [CDOM or detritus]; Siegel
and Michaels 1996). CDOM is, in turn, derived partly from
the PP that has occurred earlier in the system. The positive
correlation found between chlorophyll in the surface layer
and K
d
indicates that light attenuation also is directly
affected by the amount of phytoplankton present in the
surface layer. However, as only 30% of the interannual
variation in K
d
in this study can be explained by variation
in surface layer chlorophyll, we attribute a major part of
the variation in K
d
occurring here to be related to the
amount of dissolved organic and particulate matter and the
light attenuation caused by water itself. It has earlier been
shown for this region that less than 30% (and often much
less) of the light attenuation is caused directly by
chlorophyll (Markager et al. 2010; Timmermann et al.
2010; Krause-Jensen et al. 2012). Thus, the interannual
relationship between N-loading and K
d
is directly affected
by the amount of chlorophyll in the surface layer and
further strengthened by the dissolved organic and partic-
ulate matter that comes with runoff water from land.
PP in the PBL—The
estimates of PP made with the
assumptions used here suggest that a significant proportion
of the PP in the study area occurs in the PBL. The average
contribution of this DPP to the annual PP for the six
stations ranged from 6% to 30%. This estimate agrees well
with previous findings in the region (Richardson and
Christoffersen 1991; Richardson et al. 2000, 2003).
Furthermore, the water column PP estimates indicate that
photosynthesis in the PBL is a consistent feature during the
period extending from April to October.
The chlorophyll-specific P vs. E parameters give support
to the conclusion derived from the PP estimates that a
photosynthetically active phytoplankton population is
found in the PBL. These parameters also suggest that the
phytoplankton population in the PBL is physiologically
distinct from populations in the surface layer. Both I
k
and
P
B
are lower in samples taken in the PBL than for
max
samples in the surface layer. This suggests that light is a
limiting factor in the PBL. The obvious adaptation and
acclimatization to low light would be higher
a
values.
However, chlorophyll-specific
a
values cannot be used to
detect this as
a
and chlorophyll content per cell will co-
vary. Chl
a
concentration within a single phytoplankton
cell increases during the process of acclimatization to lower
light intensities (Cullen 1982; Richardson et al. 1983).
Therefore, the lack of difference in
a
B
between depths may
be due to higher chlorophyll contents per cell for
phytoplankton from depth 2. Unfortunately, cell counts
were not made for the deep samples so we could not
calculate
a
per cell.
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1688
Lyngsgaard et al.
here for the BSTZ has potentially important implications for
the production of oxygen in the PBL. Converting the
average DPP from this study (17% of annual production) to
oxygen equivalents (O
2
: CO
2
5
1) yields an oxygen
production in the PBL of about 89 g O
2
m
22
yr
21
. Jørgensen
and Richardson (1996) examined data from the Southern
Kattegat and estimated the oxygen demand in the bottom
layer to be 202 g O
2
m
22
yr
21
. Hansen and Bendtsen (2013)
argue that 28% of PP (, 54 g C m
22
yr
21
based on data
from this study) is remineralized in the bottom waters. This
would be equivalent to an approximate oxygen demand of
144 g O
2
m
22
yr
21
. Thus, the present study suggests that
oxygen produced by DPP in the BSTZ may compensate for a
considerable fraction of the oxygen demand in this layer.
Subsurface PP has also been shown to counteract,
although not eliminate, hypoxia in other areas where light
penetrates into the PBL (Lehrter et al. 2009; Murrell et al.
2009; Strom et al. 2010). Studies on the importance of
oxygen production below the surface layer for the
Louisiana continental shelf showed that up to 41% of the
bottom water oxygen demand may be resupplied by deep
PP, and when benthic production is included, this oxygen
contribution may be even higher (Dortch et al. 1994;
Lehrter et al. 2009).
Implications for understanding ecosystem effects of
anthropogenic nutrient loading—Nutrient
enrichment is well
known to influence both the magnitude of the production
of organic material in a system (Nixon 1995) and light
attenuation in the water column (Nielsen et al. 2002). Until
now, most studies of eutrophication in marine systems
have, therefore, focused on the influence of nutrient
enrichment on the total water column PP. The current
study suggests, however, that for the BSTZ, changes in
land-based N-loading influence the vertical distribution of
PP more than the total magnitude of PP. We see no reason
to believe that this should not be the case for other
stratified coastal systems as well.
This vertical redistribution of PP is potentially important
for the function of the system as a whole, as photosynthesis
occurring in the PBL introduces oxygen into the deeper
regions of the water column, regions that are not directly
able to exchange oxygen with the atmosphere and, thus,
ameliorate hypoxia. In addition, a change in the vertical
distribution of PP, as reported here, will likely have
consequences for the relative occurrence of different
phytoplankton species. This can potentially have conse-
quences both for food webs and for the amount of organic
material reaching bottom waters. How this vertical
redistribution of PP may affect ecosystem function
remains, however, to be quantified.
Acknowledgments
This study received funding from grants 2104-09-063212 and
2104-09-67259 from the Strategic Research Council of Denmark.
Additional support was received from the Danish National
Research Foundation via a grant to the Center for Macroecology
Evolution and Climate, University of Copenhagen, and from the
Department for Bioscience, Aarhus University. We thank
the Department for Bioscience, Aarhus University, for access to
the monitoring data and all the people that have analyzed the
The data did not allow us to differentiate between
adaptation occurring at the population level (i.e., through
changing species composition) and that occurring as
acclimatization at the cellular level. Given, however, the
very different light and nutrient conditions found in the
surface layer and PBL, respectively, it seems likely that
different species may have populated the two environ-
ments. Earlier studies have documented significant vertical
heterogeneity in the distribution of phytoplankton species
in the southern Kattegat (Mouritsen and Richardson 2003).
Further evidence that the populations found in the PBL are
photosynthetically active is found in the seasonal changes
(highest values being found during summer months, when
light intensities are greatest) noted for the P vs. E
parameters derived from populations from the PBL.
Other nitrogen sources—The
fact that we identify such a
clear relationship between the land-based N-loadings origi-
nating from Denmark and the vertical distribution of PP
occurring at six stations in the BSTZ may, initially, seem
surprising, as the BSTZ receives nutrients from several sources
(e.g., advection from surrounding seas, atmospheric deposi-
tion, and land-based runoff from several countries). Nitrogen
budgets for the area (Jørgensen et al. 2013) show that land-
based N-loading only constitute 14% of the total nitrogen
input to the region and 21% of the bioavailable N input. In
addition, only about half of the land-based loadings come
from Denmark. The explanation for the relationship recorded
may possibly be found in the fact that all of the stations in this
study are located close to the Danish coast. As a result, land-
based loading from Germany and Sweden is largely processed
prior to its reaching the sample locations. We attempted to
include the German and Swedish loadings in the regression
analyses (data not shown) but this gave poorer fits than when
we only included the Danish land-based loadings.
Nitrogen entering the BSTZ from adjacent seas consti-
tutes 58% of the bioavailable N (Jørgensen et al. 2013)
introduced to the region. However, these inputs probably
constitute a rather-constant background that leaves the
changes in land-based loading as a likely potential
candidate in generating interannual variability in PP.
Furthermore, some of those inputs may not be reaching
the surface layer during the productive season (see
arguments in Jørgensen et al. 2013).
The strong coupling between local loading and PP means
that local load reductions also can be effective in improving
the marine environment in the area. Previous studies (as
discussed in Duarte 2009) suggest that a return to the
previous condition after load reductions is not to be
expected based on data for phytoplankton biomass.
However, the results in Table 2 and Fig. 6 indicate a linear
response between PP and N-loading in the BSTZ. The
relationship found in this study between PP and N-loading
suggests that eutrophication may, indeed, be reversible and
that it can be abated by reduction in local N-loading, as
phytoplankton PP is the key process starting the cascade of
effects in marine nutrient eutrophication.
Oxygen production in the PBL—The
vertical redistribu-
tion of PP in relation to land-based N-loading demonstrated
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
Baltic primary production and N-loading
samples over the years; Morten Holtegaard Nielsen for producing
the map of the research area; and the reviewers for valuable
comments.
1689
References
¨
B
ENDTSEN
, J., K. E. G
USTAFSSON
, J. S
ODERKVIST
,
AND
J. L. S.
H
ANSEN
. 2009. Ventilation of bottom water in the North Sea–
Baltic Sea transition zone. J. Mar. Syst.
75:
138–149,
doi:10.1016/j.jmarsys.2008.08.006
B
OESCH
, D. 2002. Challenges and opportunities for science in
reducing nutrient over-enrichment of coastal ecosystems.
Estuaries
25:
886–900, doi:10.1007/BF02804914
B
ORUM
, J.,
AND
K. S
AND
-J
ENSEN
. 1996. Is total primary production
in shallow coastal marine waters stimulated by nitrogen
loading? Oikos
76:
406–410, doi:10.2307/3546213
´
C
ARSTENSEN
, J., M. S
ANCHEZ
-C
AMACHO
, C. M. D
UARTE
, D.
`
K
RAUSE
-J
ENSEN
,
AND
N. M
ARBA
. 2011. Connecting the dots:
Responses of coastal ecosystems to changing nutrient
concentrations. Environ. Sci. Technol.
45:
9122–9132,
doi:10.1021/es202351y
C
LOERN
, J. E. 2001. Our evolving conceptual model of the coastal
eutrophication problem. Mar. Ecol. Prog. Ser.
210:
223–253,
doi:10.3354/meps210223
C
ONLEY
, D. 1999. Biogeochemical nutrient cycles and nutrient
management strategies. Hydrobiologia
410:
87–96,
doi:10.1023/A:1003784504005
C
ONLEY
, D. J., J. C
ARSTENSEN
, G. Æ
RTEBJERG
, P. B. C
HRISTENSEN
,
T. D
ALSGAARD
, J. L. S. H
ANSEN
,
AND
A. B. J
OSEFSON
. 2007.
Long-term changes and impacts of hypoxia in Danish coastal
waters. Ecol. Appl.
17:
165–184, doi:10.1890/05-0766.1
———, S. M
ARKAGER
, J. H. A
NDERSEN
, T. E
LLERMAN
,
AND
L. M.
S
VENDSEN
. 2002. Coastal eutrophication and the Danish
National Aquatic Monitoring and Assessment Program.
Estuaries
25:
706–719, doi:10.1007/BF02804910
C
ULLEN
, J. J. 1982. The deep chlorophyll maximum: Comparing
vertical profiles of chlorophyll
a.
Can. J. Fish.Aquat. Sci.
39:
791–803, doi:10.1139/f82-108
D
ANISH
S
TANDARDS
, D. 1986. Vandundersøgelse 2201. Chloro-
phyll
a.
Spectrophotometric measurement in ethanol extracts.
Danish Standards. Water Examination 2201 [References
included].
D
ORTCH
, Q., N. R
ABALAIS
, R. T
URNER
,
AND
G. R
OWE
. 1994.
Respiration rates and hypoxia on the Louisiana shelf. Estuar.
Coasts
17:
862–872, doi:10.2307/1352754
D
UARTE
, C. M. 2009. Coastal eutrophication research: A new
awareness, p. 263–269.
In
J. H. Andersen and D. J. Conley
[eds.], Eutrophication in coastal ecosystems. Springer.
F
OFONOFF
, N. P. 1985. Physical properties of seawater: A new
salinity scale and equation of state for seawater. J. Geophys.
Res. Oceans
90:
3332–3342, doi:10.1029/JC090iC02p03332
G
USTAFSSON
, B. G. 2000. Time-dependent modeling of the Baltic
entrance area. 2. Water and salt exchange of the Baltic Sea.
Estuaries
23:
253–266, doi:10.2307/1352831
H
ANSEN
, J. L. S.,
AND
J. B
ENDTSEN
. 2013. Parameterisation of
oxygen dynamics in the bottom water of the Baltic Sea–North
Sea transition zone. Mar. Ecol. Prog. Ser.
481:
25–39,
doi:10.3354/meps10220
J
ØRGENSEN
, B. B.,
AND
K. R
ICHARDSON
. 1996. Eutrophication in
coastal marine ecosystems. Coast. Estuar. Stud.
52:
273.
J
ØRGENSEN
, L., S. M
ARKAGER
,
AND
M. M
AAR
. 2013. On the
importance of quantifying bioavailable nitrogen instead of
total nitrogen. Biogeochemistry
117:
1–18.
K
AAS
, H.,
AND
S. M
ARKAGER
. 1998. Technical guidlines for marine
monitoring. National Environmental Research Institute,
Denmark.
´
K
ARLSON
, B., L. E
DLER
, W. G
RANELI
, E. S
AHLSTEN
,
AND
M.
K
UYLENSTIERNA
. 1996. Subsurface chlorophyll maxima in
the Skagerrak—processes and plankton community structure.
J. Sea Res.
35:
139–158, doi:10.1016/S1385-1101(96)90742-X
K
EMP
, M. W.,
AND
W. R. B
OYNTON
. 1984. Spatial and temporal
coupling of nutrient inputs to estuarine primary production:
The role of particulate transport and decomposition. Bull.
Mar. Sci.
35:
522–535.
K
RAUSE
-J
ENSEN
, D., S. M
ARKAGER
,
AND
T. D
ALSGAARD
. 2012.
Benthic and pelagic primary production in different nutrient
regimes. Estuar. Coasts
35:
527–545, doi:10.1007/s12237-011-
9443-1
L
EHRTER
, J. C., M. C. M
URRELL
,
AND
J. C. K
URTZ
. 2009.
Interactions between freshwater input, light, and phytoplank-
ton dynamics on the Louisiana continental shelf. Cont. Shelf
Res.
29:
1861–1872, doi:10.1016/j.csr.2009.07.001
L
EWIS
, M. R., J. J. C
ULLEN
,
AND
T. P
LATT
. 1984. Relationships
between vertical mixing and photoadaptation of phytoplank-
ton: Similarity criteria. Mar. Ecol. Prog. Ser.
15:
141–149,
doi:10.3354/meps015141
L
UND
-H
ANSEN
, L. C.,
AND OTHERS
. 2008. A consistent high
primary production and chlorophyll-a maximum in a narrow
strait—effects of hydraulic control. J. Mar. Syst.
74:
395–405,
doi:10.1016/j.jmarsys.2008.03.005
M
ARKAGER
, S. 1993. Light absorption and quantum yield for
growth in five species of marine macroalgae. J. Phycol.
29:
54–63, doi:10.1111/j.1529-8817.1993.tb00279.x
———. 1998. Dark uptake of inorganic
14
C in oligotrophic
oceanic waters. J. Plankton Res.
20:
1813–1836, doi:10.1093/
plankt/20.9.1813
———, J. C
ARSTENSEN
, D. K
RAUSE
-J
ENSEN
, J. W
INDOLF
,
AND
K.
T
IMMERMANN
. 2010. Effects of increased nitrogen load on the
environment in Danish Fjords. National Environmental
Research Institute, Aarhus Univ.
———, W. V
INCENT
,
AND
E. Y. T
ANG
. 1999. Carbon fixation by
phytoplankton in high Arctic lakes: Implications of low
temperature for photosynthesis. Limnol. Oceanogr.
44:
597–607, doi:10.4319/lo.1999.44.3.0597
———,
AND
W. F. V
INCENT
. 2001. Light absorption by
phytoplankton: Development of a matching parameter for
algal photosynthesis under different spectral regimes. J.
Plankton Res.
23:
1373–1384, doi:10.1093/plankt/23.12.1373
M
OURITSEN
, L. T.,
AND
K. R
ICHARDSON
. 2003. Vertical microscale
patchiness in nano- and microplankton distributions in a
stratified estuary. J. Plankton Res.
25:
783–797, doi:10.1093/
plankt/25.7.783
M
URRELL
, M. C., J. G. C
AMPBELL
, J. D. H
AGY
, III,
AND
J. M.
C
AFFREY
. 2009. Effects of irradiance on benthic and water
column processes in a Gulf of Mexico estuary: Pensacola Bay,
Florida, USA. Estuar. Coast. Shelf Sci.
81:
501–512,
doi:10.1016/j.ecss.2008.12.002
N
IELSEN
, E. S. 1952. The use of radio-active carbon (
14
C) for
measuring organic production in the sea. J. Cons. int. Explor.
Mer
18:
117–140, doi:10.1093/icesjms/18.2.117
N
IELSEN
, S., K. S
AND
-J
ENSEN
, J. B
ORUM
,
AND
O. G
EERTZ
-H
ANSEN
.
2002. Depth colonization of eelgrass (Zostera
marina)
and
macroalgae as determined by water transparency in Danish
coastal waters. Estuaries
25:
1025–1032, doi:10.1007/
BF02691349
N
IXON
, S. W. 1995. Coastal marine eutrophication: A definition,
social causes, and future concerns. Ophelia
41:
199–219.
˚
P
ETERSEN
, D. L. J.,
AND
M. H
JORTH
. 2010. Marine omrader 2009.
NOVANA. State and trends in the environment and nature
quality. Faglig rapport fra DMU 800 Danmarks Miljøunder-
søgelser, Aarhus Univ., Roskilde. NERI Technical Report.
MOF, Alm.del - 2015-16 - Endeligt svar på spørgsmål 866: Spm. om professor Stiig Markager materiale om danske kvælstoftilførslers indflydelse på vandets klarhed i Kattegat-Bælthavet m.m. fremlagt under høringen den 23. februar 2016, til miljø- og fødevareministeren
1690
Lyngsgaard et al.
T
IMMERMANN
, K., S. M
ARKAGER
,
AND
K. E. G
USTAFSSON
. 2010.
Streams or open sea? Tracing sources and effects of nutrient
loadings in a shallow estuary with a 3D hydrodynamic–
ecological model. J. Mar. Syst.
82:
111–121, doi:10.1016/
j.jmarsys.2010.04.008
W
EBB
, W. L., M. N
EWTON
,
AND
D. S
TARR
. 1974. Carbon dioxide
exchange of
Alnus rubra.
Oecologia
17:
281–291, doi:10.1007/
BF00345747
W
INDOLF
, J., G. B
LICHER
-M
ATHIESEN
, J. C
ARSTENSEN
,
AND
B.
K
RONVANG
. 2012. Changes in nitrogen loads to estuaries
following implementation of governmental action plans in
Denmark: A paired catchment and estuary approach for
analysing regional responses. Environ. Sci. Policy
24:
24–33,
doi:10.1016/j.envsci.2012.08.009
———, H. T
HODSEN
, L. T
ROLDBORG
, S. E. L
ARSEN
, J. B
ØGESTRAND
,
N. B. O
VESEN
,
AND
B. K
RONVANG
. 2011. A distributed modeling
system for simulation of monthly runoff and nitrogen sources,
loads and sinks for ungauged catchments in Denmark. J.
Environ. Monit.
13:
2645–2658, doi:10.1039/c1em10139k
R
ICHARDSON
, K., J. B
EARDALL
,
AND
J. A. R
AVEN
. 1983. Adaptation
of unicellular algae to irradiance: An analysis of strategies.
New Phytol.
93:
157–191, doi:10.1111/j.1469-8137.1983.
tb03422.x
———,
AND
A. C
HRISTOFFERSEN
. 1991. Seasonal distribution and
production of phytoplankton in the Southern Kattegat. Mar.
Ecol. Prog. Ser.
78:
217–227, doi:10.3354/meps078217
———, B. R
ASMUSSEN
, T. B
UNK
,
AND
L. T. M
OURITSEN
. 2003.
Multiple subsurface phytoplankton blooms occurring simul-
taneously in the Skagerrak. J. Plankton Res.
25:
799–813,
doi:10.1093/plankt/25.7.799
———, A. W. V
ISSER
,
AND
F. B. P
EDERSEN
. 2000. Subsurface
phytoplankton blooms fuel pelagic production in the North Sea.
J. Plankton Res.
22:
1663–1671, doi:10.1093/plankt/22.9.1663
S
AVCHUK
, O. P. 2005. Resolving the Baltic Sea into seven
subbasins: N and P budgets for 1991–1999. J. Mar. Syst.
56:
1–15, doi:10.1016/j.jmarsys.2004.08.005
S
IEGEL
, D. A.,
AND
A. F. M
ICHAELS
. 1996. Quantification of non-
algal light attenuation in the Sargasso Sea: Implications for
biogeochemistry and remote sensing. Deep-Sea Res. II
43:
321–345, doi:10.1016/0967-0645(96)00088-4
S
MITH
, V. 2003. Eutrophication of freshwater and coastal marine
ecosystems a global problem. Environ. Sci. Pollut. Res.
10:
126–139, doi:10.1065/espr2002.12.142
S
TRICKLAND
, J. D. H.,
AND
T. R. P
ARSONS
. 1972. A practical handbook
of seawater analysis. Bull. Fish. Res. Bd. Can 167.
S
TROM
, S. L., E. L. M
ACRI
,
AND
K. A. F
REDRICKSON
. 2010. Light
limitation of summer primary production in the coastal Gulf
of Alaska: Physiological and environmental causes. Mar.
Ecol. Prog. Ser.
402:
45–57, doi:10.3354/meps08456
Associate editor: Anthony W. D. Larkum
Received: 07 November 2013
Accepted: 23 May 2014
Amended: 28 May 2014